Abstract
Oligomeric proteins are central to cellular life and the duplication and divergence of their genes is a key driver of evolutionary innovations. The duplication of a gene coding for an oligomeric protein has numerous possible outcomes, which motivates questions on the relationship between structural and functional divergence. How do protein oligomeric states diversify after gene duplication? In the simple case of duplication of a homo-oligomeric protein gene, what properties can influence the fate of descendant paralogs toward forming independent homomers or maintaining their interaction as a complex? Furthermore, how are functional innovations associated with the diversification of oligomeric states? Here, we review recent literature and present specific examples in an attempt to illustrate and answer these questions.
Current Opinion in Genetics & Development 2022, 76:101966
This review comes from a themed issue on Evolutionary Genetics
Edited by Christian Landry and Gianni Liti
For complete overview of the section, please refer to the article collection, “Evolutionary Genetics”
Available online 22th August 2022
https://doi.org/10.1016/j.gde.2022.101966
0959-437X/© 2022 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Introduction
One of the earliest accounts of gene duplication appeared in the 1930s, when doubling of a chromosomal band in Drosophila melanogaster was associated with extreme reduction in eye size [1]. Since then, this phenomenon has been recognized as a ubiquitous source of functional novelty across the Tree of Life 2, 3, 4. Indeed, duplicated genes (typically termed as paralogs [5]) initially retain the overall sequence and structural features of the ancestral single-copy gene, but over time, they also diverge, either symmetrically [6] or asymmetrically 7, 8, 9, that is, at similar or different rates, and can bring about functional innovations.
Functional innovations are intimately associated with changes in protein structure. In this context, it is important to consider the tertiary and quaternary structure of paralogous proteins when examining the evolution of their function. For example, after duplication, a protein may acquire a novel targeting sequence and perform the ancestral function in a new subcellular compartment [10]. Alternatively, sequence changes in paralogs by point-mutations or by insertion-deletions can lead to functional innovations 11, 12, 13•, 14••, 15, 16, 17, 18, 19, 20, 21• such as subdivision of the ancestral function among the duplicated copies [11], new enzymatic activities 12, 13•, regulatory modes 12, 14••, or interaction partners 15, 16. In this review, we survey and classify how changes in quaternary structure relate to the functional diversification of paralogous proteins. First, we present how gene duplication can diversify protein oligomeric states in the context of homomers. Second, we examine molecular and regulatory changes associated with such diversification. Finally, we illustrate how functional innovations can be coupled to changes in oligomeric state.
Gene duplication can drive the emergence of new protein oligomeric states
The duplication of a gene opens numerous possible routes of functional innovation. Once fixed, the functional redundancy conferred by the two copies relaxes the selection pressure on them, allowing the diversification of their sequence, structure, oligomeric state, and in turn, function. A homomeric protein is perhaps the simplest model system to understand these events (Figure 1a). In principle, right after duplication of a gene encoding a homomeric protein, a statistical mixture of homo- and heteromeric complexes would form (Mixed). Upon further divergence, various scenarios may occur 22••, 23, 24••, 25. The statistical mixture may continue to exist 23, 26. Alternatively, the two copies might lose the capacity to cross-interact, resulting in two independent homomers (Obligate homo); or, they might lose the capacity to self-interact, resulting in a single heteromeric complex (Obligate hetero). The interaction pattern might also diverge asymmetrically, in which one copy retains the ancestral homomeric state, whereas the other evolves heteromeric interactions with new partners (Hetero others). The relative frequency of these four fates varies, depending on the organism and the symmetry of the homomer [22].
Figure 1.
Evolution of oligomeric state among homomeric proteins upon gene duplication. (a) Duplication of a gene encoding a homomeric protein. Before divergence, interfaces are compatible, so a mixture of homo- and heteromeric complexes can coexist. As the two copies diverge, three outcomes may follow: (i) each duplicate self-interacts, which we refer to as “Obligate homomers”. (ii) Both duplicates can interact to form a heteromeric complex that we refer to as “Obligate heteromer”. (iii) The duplicates follow different paths, with, for example, one copy remaining a homomer, while the other gains a new interaction. (b) The relative frequency of the four different outcomes as observed in an analysis of E. coli and S. cerevisiae’s interaction networks [22]. Homomeric interactions are dominant in E. coli, whereas cross-reacting paralogs are dominant in S. cerevisiae. (c) After duplication of a gene coding for a homomeric protein, the two paralogs can fix as an Obligate heteromeric complex. Following such an event, further duplications events may expand the family. (d) Gene duplication of ring-like homomers often involves the paralogous copies co-assembling into the same ring, making the ring heteromeric.
The retention of homomeric interactions is the dominant fate in bacteria, where Obligate homomers account for more than half of the examined cases (Figure 1b). In this scenario, paralogous copies tend to be substantially different in both sequence and primary functions [27]. Though the underlying principles await discovery, given that in bacteria horizontal gene transfers are more than ~50 times more frequent than gene duplications [28], many of these Obligate homomers may not be bona fide paralogs. Instead, they might include a horizontally acquired gene of the same family. In this case, the homologous proteins may significantly differ in sequence and would not cross-interact [29]. In contrast, horizontal transfers are rare in eukaryotes and nearly two-thirds of their homomeric gene duplications yield cross-interacting paralogs (Obligate hetero+Mixed dominate, Figure 1b). When an ancestral homomer becomes an Obligate heteromer, the primary function before and after the duplication tends not to change. Phosphofructokinase (PFK, E.C. 2.7.1.11) [30] forms a homo-octamer in fission yeast. In the lineage of baker’s yeast, it underwent a gene duplication event resulting in a hetero-octameric PFK1/2 complex that retained the ancestral enzymatic activity (E.C. 2.7.1.11) [31]. Though the primary function is conserved, the sequence and structure of each copy can evolve toward the emergence or fine-tuning of secondary function(s) 3, 32••, or asymmetric binding of partners [23]. This route underlies the emergence of regulatory modes in many eukaryotic enzymes typically absent among their prokaryote orthologs [3].
We saw how gene duplication can yield an Obligate heteromer. Interestingly, once two paralogs are fixed as a heteromeric complex, their further duplications might generate an array of heteromeric complexes (Figure 1c). This phenomenon underlies the expansion of many generalist enzyme families (those that work on a broad range of substrates), including transporters, chaperones, or endonucleases. Perhaps, the most intriguing example is the HSP70 chaperone family. A gene duplication in the HSP70 family during the emergence of eukaryotes diverged into the HSP110 cochaperone that lost the canonical chaperone activity, but specialized to become an ATP-exchange factor for HSP70 [33]. This functional collaboration between HSP70 and HSP110 has been retained in their successive duplications, and with the ever-increasing demand for chaperones in eukaryotes, has rendered an array of HSP70–HSP110 pairs that work in different conditions, subcellular compartments, tissues, and also on different types of substrates 33, 34.
For ring-like homomers, paralogous copies typically, but not always, sequester in the same ring, thus making the ring heteromeric (Obligate hetero is the dominant fate, Figure 1d). Examples include ATPase rings 35, 36, the chaperonin TCP complex [37], the RNA exosome PH ring [38], Lsm/Sm homomers [39], and the V0 ring of V-ATPase proton pump [40] that have become heteromeric in eukaryotes following one or several rounds of gene duplication. It is interesting to note that for numerous complexes with ring-like (cyclic) symmetry, the patterns of gene duplications appear to be discrete. For example, the proteasome alpha ring (C7 symmetry) is formed by one protein in archaea and seven paralogous proteins in eukaryotes, but intermediate numbers (e.g., 2, 3, 4, 5, or 6 paralogous proteins) have not been observed, presumably due to symmetry constraints 23, 41••. Interestingly, after the proteasomal alpha and beta rings were fully heteromeric, whole-genome duplication(s) at the root of jawed vertebrates gave birth to the immunoproteasome — differing from the more common “constitutive proteasome” by a few paralogous subunits and specialized to operate during oxidative stress [42].
Molecular and regulatory changes associated with interaction divergence
Immediately after a duplication event, the interfaces of both duplicates are identical and therefore compatible with preduplication interaction modes. As the duplicates diverge, at least four distinct mechanisms can drive interaction divergence between the paralogous copies. First, divergence can emerge from different subcellular localizations and/or different modes of expression [43] of the two paralogs (e.g., at different stages of the cell cycle or in different tissues, Figure 2a). Until their interfaces diverge by random genetic drift, such recently duplicated Obligate homomers can still cross-interact even though they may not encounter each other in vivo. Examples include cytoplasmic and mitochondrial homomeric thioredoxin reductases in yeast (TRR1 and TRR2) that emerged during the yeast whole-genome duplication [44]. These two paralogs are 85% identical in sequence and were observed to cross-interact in vitro [45].
Figure 2.
Origins of interaction divergence between paralogs. (a) Incompatibility may emerge from different subcellular localization of the paralogous copies, such as the case of yeast Obligate homomers TRR1 (PDB code 3DX8) and TRR2 (UniProt code P38816). (b) Insertions and deletions at the interface can drive incompatibility, as in the case of yeast MAS1/2 heterodimer where an insertion (Ile42–Thr45) in MAS1 appears to block its self-interaction. In another example, two insertions, Gln211–Asp231 and Asp299–Thr365, appear to hinder filament formation of ARP4, which instead heteromerizes with its paralog ACT1. Here, the structure depicts an ACT1 filament (white, PDB code 6BNO) onto which ARP4 (purple, PDB code 5NBM) was superposed. (c) The gain/loss of the domain mediating the ancestral homomeric interaction is another mechanism for incompatibility. Such a scenario is seen in Obligate homomers GAL10 (PDB code 1Z45) and YNR071C (PDB code 1YGA). (d) Accumulation of amino acid substitutions can alter the interface specificity and bring about incompatibility. This scenario occurs in the yeast heteromer IDH1/2 (PDB code 3BLV). We note that the IDH complex shows A4B4 stoichiometry and only half of the complex (A2B2) is shown here for simplicity. The physiological relevance of the quaternary structures highlighted in this figure was inferred based on annotations from the QSBIO.org database [83].
Second, insertions/deletions (InDels) at the interface can introduce steric hindrance and create interface incompatibility (Figure 2b) 46, 47. For example, MAS1 and MAS2 assemble into the Obligate heterocomplex yeast mitochondrial peptidase. The two genes encoding these proteins share a common bacterial ancestor, which forms a homodimer (M16 family peptidase) [22]. The heteromeric interface of the MAS1/2 complex shares a similar structure to that of the bacterial M16 peptidases and an insertion in MAS1 (Ile42–Thr45) would hinder its homodimerization. In another example, yeast actin (ACT1) forms a homomeric filament [48]. Its paralog, actin-related protein 4 (ARP4), does not form homomeric filaments and instead heterodimerizes with ACT1. The ARP4–ACT1 heteromer is a component of chromatin remodeling complexes [49]. As depicted in Figure 2b, two insertions, Gln211–Asp231 and Asp299–Thr365 appear incompatible with self-interaction interfaces found in actin filaments.
Third, divergence can originate in the gain or loss of a domain mediating or preventing an interaction (Figure 2c). An interesting case is that of yeast paralogs GAL10 and YNR071C. As annotated in Pfam [50], GAL10 comprises an N-terminal GDP-mannose-4,6-dehydratase domain. This domain mediates the homomeric interaction and also contacts the C-terminal Aldose-1-epimerase domain [51]. In contrast, YNR071C has lost the N-terminal domain according to Pfam annotations, and comprises only an Aldose-1-epimerase domain, which mediates the homomeric interaction seen in the crystal structure [51]. The cross-interaction between the two paralogs via the Aldose-1-epimerase domain is hindered by the GDP-mannose-4,6-dehydratase domain in GAL10.
Last, the accumulation of amino acid substitutions at the interface can also bring about incompatibility and divergence (Figure 2d). Indeed, a few mutations can create or abolish interaction interfaces 52••, 53, 54. The mitochondrial NAD-dependent isocitrate dehydrogenase complex in yeast represents such an example. It is an Obligate heteromer of IDH1 and IDH2 that are 42% identical in sequence but comprise no domain gain, loss, or InDels at the interface. IDH1 and IDH2 do not exhibit homomeric interactions in vitro [22], so this incompatibility appears to stem from the accumulation of amino acid substitutions at their interface.
Functional innovations are coupled with the diversification of oligomeric states
Functional innovations after gene duplication can be coupled with the divergence of the oligomeric state. For example, Obligate homomers, at least in principle, evolve independently without the constraints of paralog interference and can therefore diverge in function more than heteromers of paralogs 55, 56. Functional divergence of Obligate homomers can be classified into two categories. In the first category, the primary functions of the duplicated copies do not change but one or both become specialized to operate in different conditions (i.e., subfunctionalization, Figure 3a). For example, the duplicated copy might diverge and perform the same function in a different subcellular compartment. Yeast ALT1 and ALT2 represent such an example: both synthesize pyruvate from L-alanine, but the former localizes in mitochondria and the latter in the cytoplasm 57, 58. Subfunctionalization can also be associated with different expression profiles, as in the case of Escherichia coli LYSS and LYSU. Both catalyze lysine–tRNA loading, but the former is constitutively expressed, whereas the latter is heat-induced [59]. In the second category, the primary function itself can diverge (neofunctionalization, Figure 3b). For enzymes, the divergence of primary function could mean changes in substrate specificity, whereby both paralogs perform the same chemistry on different substrates [21]. Examples include E. coli LACA, which catalyzes the acetylation of galactosyl units [60]; its paralog, MAA, catalyzes the acetylation of glucosyl units [61]. In other cases, one copy may lose the enzyme activity altogether. Examples include yeast Obligate homomers LPD1 and IRC15 (duplicated during the whole-genome duplication [62]). The former is a lipoamide dehydrogenase enzyme [63], while the latter has lost its enzymatic activity and regulates microtubule dynamics [64].
Figure 3.
Evolution of the oligomeric state in connection to functional innovations after gene duplication. (a) The duplication of a homomer yields two independent Obligate homomers sharing the same primary function but different localizations: yeast ALT1 (mitochondrial) and ALT2 (cytosolic). (b) The duplication of a homomer yields two independent Obligate homomers exhibiting different primary functions: E. coli LACA and MAA. The former is specific for acetylation of galactosyl units [60], while the latter catalyzes acetylation of glucosyl units [61]. (c)Obligate heteromers with independent active sites can retain the ancestral function (white star), or maintain the function through additional subunits. In the case of yeast IDH1/2 heteromer, IDH1 has lost its catalytic activity and became a regulatory subunit [65]. In the case of the exosome, all paralogs forming the PH ring lost the RNAse activity and a new subunit assumed that function. (d) Among Obligate heteromers where the active site is contributed by both subunits, both subunits are expected to maintain their function. In this case, regulatory activities may emerge in a separate domain, as in the case of yeast PFK1/2 heteromer [66]. In a different scenario, an Obligate heteromer such as the yeast SPOTS complex recruits new regulatory subunits.
For Obligate heteromers, there is little scope for the divergence of the primary function because before and after the duplication there is only one complex carrying out the ancestral function. However, since two proteins are now mediating the primary function, selection pressure may relax and allow the emergence of secondary functions such as allosteric regulation [3]. Such allosteric regulation can emerge in different ways, depending on the catalytic (in)dependence of the individual subunits.
A homomeric enzyme can exhibit structurally independent catalytic sites in each subunit, or catalytic sites may be shared across multiple subunits. In the former case, once there is an Obligate heteromeric complex after gene duplication, regulatory activities can emerge in the following ways. First, one paralog might lose the catalytic activity and instead become a regulatory subunit of the other (Figure 3c), as previously discussed with the HSP70–HSP110 pair. Another example is the yeast mitochondrial NAD-dependent IDH1/2 complex (duplicated and diverged from a bacterial homomeric IDH complex [22]), in which IDH1 lost the catalytic activity and became a regulatory subunit of IDH2 [65]. In a more extreme case, all paralogous subunits may lose the catalytic activity and an external subunit may be recruited for catalysis (Figure 3c). Such a scenario occured with the bacterial homomeric RNase PH ring. After several rounds of gene duplications, it became heteromeric in eukaryotes (RNA exosome) and in that process all paralogous copies lost their catalytic activities. The exonuclease function in contemporary eukaryotic RNA exosome is mediated by a newly recruited, eukaryote-specific Rrp44 subunit [38].
For homomeric enzymes with catalytic sites contributed by multiple subunits, the scope of evolving allosteric regulation is more limited because each subunit is constrained to retain the primary function [85]. In these cases, regulatory activities can emerge within each subunit (Figure 3d). A curious example is yeast ATP-dependent 6-phosphofructokinase, an Obligate heterocomplex of PFK1 and PFK2. In this complex, the interacting N-terminal domains of the two subunits share the catalytic sites, whereas the C-terminal domains have evolved allosteric regulation [66]. Alternatively, the newly emerged heteromeric enzyme can act as a catalytic scaffold that recruits additional regulatory subunits (Figure 3d). Examples include the SPOTS complex in yeast, composed of catalytic subunits LCB1 and LCB2 (that share the serine palmitoyltransferase catalytic sites) and regulatory subunits ORM1, ORM2, SCA1, and TSC3 67, 68.
Conclusion
The transition of protein oligomeric states upon gene duplication, once only studied and appreciated for a few prokaryotic homomeric ring-like complexes 35, 36, 37, 38, 39, can now be examined as a genome-scale phenomenon. A remarkable trend of this process is that it has transformed homomer-dominant prokaryotic proteomes into heteromer-dominant eukaryotic proteomes 22••, 69, 70. The spectrum of protein oligomeric-state diversification is wide, and each mode of diversification — combined with the symmetry of the oligomeric state and the catalytic independence of the subunits — constraints functional diversification differently. Conspicuous functional innovations, such as loss of enzyme activity or alteration of substrate specificity, are typically observed when the paralogous pairs form independent complexes and do not cross-react. Comparatively more subtle functional innovations, such as the emergence of allosteric regulation, are typically observed when the paralogous pairs co-assemble into the same complex. Whether or not, two paralogs, descendants of an ancestral homomer, would form a heteromeric complex, depends on post-duplication changes at the interface. These molecular changes in paralogous enzymes have likely played an important role in the emergence of eukaryotic cells with their many subcellular compartments.
Future directions
We saw that specific evolutionary processes govern the interactions and functional divergence of gene duplicates. While this review focuses on specific examples, we still lack a systematic genome-scale understanding of these processes within and between species. In particular, we highlight below four questions. Addressing these will be important to gain fundamental knowledge, but also for understanding the association between gene duplication and various complex diseases 71, 72, 73, including Down syndrome [74], Alagille syndrome [75], and cancer [73].
First, on what timescales do evolutionary events accumulate to drive structural and functional divergence? Timescales can vary broadly in different cases, and work will be needed to identify determinants associated with different rates. This point is also tied to a second question: to what extent is divergence driven by adaptive versus neutral processes 76, 77, 78? In the former case, we can anticipate signatures of positive selection and a faster sequence divergence than in the latter. Interestingly, neutral drift is not necessarily expected to disrupt homomeric interactions [79]. In contrast, it can disrupt interactions between paralogs more frequently by altering co-expression and co-localization tendencies. Therefore, we expect that the retention of interactions between paralogs will more often be adaptive than the retention of homomeric interactions [22].
The expected balance of adaptive versus neutral evolution is itself tied to a third question: do functional innovations tend to exist pre duplication, or occur post duplication? Promiscuous activities can emerge in enzymes by neutral drift 76, 77 and come to evolve independently in a duplicate. In this context, the new function may predate the gene duplication. Alternatively, the new function could evolve in the newly formed complex post duplication.
Finally, which of the two fates, Obligate homo or Obligate hetero emerges more frequently and why? Obligate homomeric pairs dominate bacterial proteomes, whereas Obligate heteromers dominate in eukaryotes [22]. Since a few mutations 52••, 53, 54, 84 or a single insertion–deletion 47, 80 can alter the interaction specificity of protein interfaces, the sole effect of mutations on structure is unlikely to explain this dichotomy. Therefore, other factors will need to be investigated.
More generally, addressing questions will require comprehensive information on protein structure and assembly. In that respect, recent structure and interaction modeling methods based on deep learning 81, 82 will be instrumental to perform such analyses on a phylogenetic scale.
CRediT authorship contribution statement
Saurav Mallik: Writing – original draft, visualization, Emmanuel D Levy: Writing – review and editing, visualization.
Conflict of interest
The authors declare that no conflict of interest exists.
Acknowledgments
S.M. dedicates this article to a friend, colleague, and mentor Dan S. Tawfik, who supervised part of this work. S.M. was supported by a VATAT postdoctoral fellowship, provided by Israel’s Council for Higher Education Planning and Budgeting Committee. This work was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement No. 819318), by the Israel Science Foundation (grant no. 1452/18), by a research grant from A.-M. Boucher, by research grants from the Estelle Funk Foundation, the Estate of Fannie Sherr, the Estate of Albert Delighter, the Merle S. Cahn Foundation, Mrs. Mildred S. Gosden, the Estate of Elizabeth Wachsman, the Arnold Bortman Family Foundation.
Contributor Information
Saurav Mallik, Email: saurav.mallik@weizmann.ac.il.
Emmanuel D Levy, Email: emmanuel.levy@weizmann.ac.il.
References and recommended reading
Papers of particular interest, published within the period of review, have been highlighted as:
-
•
of special interest
-
••
of outstanding interest.
- 1.Bridges C.B. The bar “gene” a duplication. Science. 1936;83:210–211. doi: 10.1126/science.83.2148.210. [DOI] [PubMed] [Google Scholar]
- 2.Ohno S. Evolution by Gene Duplication. Springer Berlin; 1970. [DOI] [Google Scholar]
- 3.Conant G.C., Wolfe K.H. Turning a hobby into a job: how duplicated genes find new functions. Nat Rev Genet. 2008;9:938–950. doi: 10.1038/nrg2482. [DOI] [PubMed] [Google Scholar]
- 4.Zhang J. Evolution by gene duplication: an update. Trends Ecol Evol. 2003;18:292–298. [Google Scholar]
- 5.Jensen R.A. Orthologs and paralogs – we need to get it right. Genome Biol. 2001;2 doi: 10.1186/gb-2001-2-8-interactions1002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Kondrashov F.A., Rogozin I.B., Wolf Y.I., Koonin E.V. Selection in the evolution of gene duplications. Genome Biol. 2002;3 doi: 10.1186/gb-2002-3-2-research0008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Holland P.W.H., Marlétaz F., Maeso I., Dunwell T.L., Paps J. New genes from old: asymmetric divergence of gene duplicates and the evolution of development. Philos Trans R Soc Lond B Biol Sci. 2017;372(1713) doi: 10.1098/rstb.2015.0480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Wagner A. Asymmetric functional divergence of duplicate genes in yeast. Mol Biol Evol. 2002;19:1760–1768. doi: 10.1093/oxfordjournals.molbev.a003998. [DOI] [PubMed] [Google Scholar]
- 9.Scannell D.R., Wolfe K.H. A burst of protein sequence evolution and a prolonged period of asymmetric evolution follow gene duplication in yeast. Genome Res. 2008;18:137–147. doi: 10.1101/gr.6341207. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Marques A.C., Vinckenbosch N., Brawand D., Kaessmann H. Functional diversification of duplicate genes through subcellular adaptation of encoded proteins. Genome Biol. 2008;9 doi: 10.1186/gb-2008-9-3-r54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Lynch M., Force A. The probability of duplicate gene preservation by subfunctionalization. Genetics. 2000;154:459–473. doi: 10.1093/genetics/154.1.459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Hittinger C.T., Carroll S.B. Gene duplication and the adaptive evolution of a classic genetic switch. Nature. 2007;449:677–681. doi: 10.1038/nature06151. [DOI] [PubMed] [Google Scholar]
- 13•.Cople S.D. Evolution of new enzymes by gene duplication and divergence. FEBS J. 2020;287:1262–1283. doi: 10.1111/febs.15299. [DOI] [PMC free article] [PubMed] [Google Scholar]; Describes how new enzyme activities can emerge by gene duplication of an ancestral promiscuous enzyme and by successive fixation of mutations in the two copies.
- 14••.Jayaraman V., Lee D.J., Elad N., Vimer S., Sharon M., Fraser J.S., Tawfik D.S. A counter-enzyme complex regulates glutamate metabolism in Bacillus subtilis. Nat Chem Biol. 2022;18:161–170. doi: 10.1038/s41589-021-00919-y. [DOI] [PMC free article] [PubMed] [Google Scholar]; An intriguing study of a counter-enzyme complex (i.e. an assembly of metabolic enzymes catalyzing opposite rather than sequential reactions) in Bacillus involving a recently duplicated pair of paralogs.
- 15.Diss G., Gagnon-Arsenault I., Dion-Coté A.-M., Vignaud H., Ascencio D.I., Berger C.M., Landry C.R. Gene duplication can impart fragility, not robustness, in the yeast protein interaction network. Science. 2017;355:630–634. doi: 10.1126/science.aai7685. [DOI] [PubMed] [Google Scholar]
- 16.Veron A.S., Kaufmann K., Bornberg-Bauer E. Evidence of interaction network evolution by whole-genome duplications: a case study in MADS-box proteins. Mol Biol Evol. 2007;24:670–678. doi: 10.1093/molbev/msl197. [DOI] [PubMed] [Google Scholar]
- 17.Latham J.A., Chen D., Allen K.N., Dunaway-Mariano D. Divergence of substrate specificity and function in the Escherichia coli hotdog-fold thioesterase paralogs YdiI and YbdB. Biochemistry. 2014;53:4775–4787. doi: 10.1021/bi500333m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Shi D., Yu X., Cabrera-Luque J., Chen T.Y., Roth L., Morizono H., Allewell N.M., Tuchman M. A single mutation in the active site swaps the substrate specificity of N-acetyl-L-ornithine transcarbamylase and N-succinyl-L-ornithine transcarbamylase. Protein Sci. 2007;16:1689–1699. doi: 10.1110/ps.072919907. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.VanderSluis B., Bellay J., Musso G., Costanzo M., Papp B., Vizeacoumar F.J., Baryshnikova A., Andrews B., Boone C., Myers C.L. Genetic interactions reveal the evolutionary trajectories of duplicate genes. Mol Syst Biol. 2010;6 doi: 10.1038/msb.2010.82. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Pascarelli S., Laurino P. Inter-paralog amino acid inversion events in large phylogenies of duplicated proteins. PLoS Comput Biol. 2022;18 doi: 10.1371/journal.pcbi.1010016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21•.Yang G., Miton C.M., Tokuriki N. A mechanistic view of enzyme evolution. Protein Sci. 2020;29:1724–1747. doi: 10.1002/pro.3901. [DOI] [PMC free article] [PubMed] [Google Scholar]; In the light of several examples, the authors discuss that after gene duplication of an ancestral enzyme with a main function and a promiscuous secondary function, one of the duplicated copies might accumulate mutations enhancing the ancestral secondary function, while the other retains the primary function.
- 22••.Mallik S., Tawfik D.S. Determining the interaction status and evolutionary fate of duplicated homomeric proteins. PLoS Comput Biol. 2020;16 doi: 10.1371/journal.pcbi.1008145. [DOI] [PMC free article] [PubMed] [Google Scholar]; A systematic analysis of the fate of gene duplication of homomeric proteins in bacteria and eukaryotes. This study also depicts that a substantial fraction of prokaryotic homomeric protein families has become heteromeric in eukaryotes by gene duplications.
- 23.Pereira-Leal J.B., Levy E.D., Kamp C., Teichmann S.A. Evolution of protein complexes by duplication of homomeric interactions. Genome Biol. 2007;8 doi: 10.1186/gb-2007-8-4-r51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24••.Marchant A., Cisneros A.F., Dubé A.K., Gagnon-Arsenault I., Ascencio D., Jain H., Aubé S., Eberlein C., Evans-Yamamoto D., Yachie N., et al. The role of structural pleiotropy and regulatory evolution in the retention of heteromers of paralogs. Elife. 2019;8 doi: 10.7554/eLife.46754. [DOI] [PMC free article] [PubMed] [Google Scholar]; Using in silico evolution, the authors showed that upon gene duplication of a homomer, heteromerization of the two descendant paralogs could be preserved indirectly due to the selection for the maintenance of the ancestral homomer.
- 25.Hochberg G.K.A., Shepherd D.A., Marklund E.G., Santhanagoplan I., Degiacomi M.T., Laganowsky A., Allison T.M., Basha E., Marty M.T., Galpin M.R., et al. Structural principles that enable oligomeric small heat-shock protein paralogs to evolve distinct functions. Science. 2018;359:930–935. doi: 10.1126/science.aam7229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Maslov S., Sneppen K., Eriksen K.A., Yan K.-K. Upstream plasticity and downstream robustness in evolution of molecular networks. BMC Evol Biol. 2004;4 doi: 10.1186/1471-2148-4-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27•.Dandage R., Landry C.R. Paralog dependency indirectly affects the robustness of human cells. Mol Syst Biol. 2019;15 doi: 10.15252/msb.20198871. [DOI] [PMC free article] [PubMed] [Google Scholar]; The authors conducted a systematic analysis of human heteromeric paralogous pairs by investigating the fitness effect of single gene knockouts by CRISPR-Cas9 in several human cell lines; they concluded that the functional co-dependency of heteromeric paralogs accounts for the deleterious nature of the knockout of either copy.
- 28•.Tria F.D.K., Martin W.F. Gene duplications are at least 50 times less frequent than gene transfers in prokaryotic genomes. Genome Biol Evol. 2021;13 doi: 10.1093/gbe/evab224. [DOI] [PMC free article] [PubMed] [Google Scholar]; The authors examined recent horizontal gene transfers among distantly related bacterial species; the results depict that horizontal gene transfers are 50–100 times more frequent than gene duplications in bacteria.
- 29.Kachroo A.H., Laurent J.M., Yellman C.M., Meyer A.G., Wilke C.O., Marcotte E.M. Evolution. Systematic humanization of yeast genes reveals conserved functions and genetic modularity. Science. 2015;348:921–925. doi: 10.1126/science.aaa0769. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Reuter R., Naumann M., Bär J., Haferburg D., Kopperschläger G. Purification, molecular and kinetic characterization of phosphofructokinase-1 from the yeast Schizosaccharomyces pombe: evidence for an unusual subunit composition. Yeast. 2000;16:1273–1285. doi: 10.1002/1097-0061(200010)16:14<1273::AID-YEA621>3.0.CO;2-#. [DOI] [PubMed] [Google Scholar]
- 31.Banaszak K., Mechin I., Obmolova G., Oldham M., Chang S.H., Ruiz T., Radermacher M., Kopperschläger G., Rypniewski W. The crystal structures of eukaryotic phosphofructokinases from baker’s yeast and rabbit skeletal muscle. J Mol Biol. 2011;407:284–297. doi: 10.1016/j.jmb.2011.01.019. [DOI] [PubMed] [Google Scholar]
- 32••.Pillai A.S., Chandler S.A., Liu Y., Signore A.V., Cortez-Romero C.R., Benesch J.L.P., Laganowsky A., Storz J.F., Hochberg G.K.A., Thornton J.W. Origin of complexity in haemoglobin evolution. Nature. 2020;581:480–485. doi: 10.1038/s41586-020-2292-y. [DOI] [PMC free article] [PubMed] [Google Scholar]; The authors reconstructed the evolutionary origin of modern hetero-tetrameric haemoglobin from an ancestral homodimeric oxygen-binding protein.
- 33.Rebeaud M.E., Mallik S., Goloubinoff P., Tawfik D.S. On the evolution of chaperones and cochaperones and the expansion of proteomes across the Tree of Life. Proc Natl Acad Sci USA. 2021;118 doi: 10.1073/pnas.2020885118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Gong Y., Kakihara Y., Krogan N., Greenblatt J., Emili A., Zhang Z., Houry W.A. An atlas of chaperone-protein interactions in Saccharomyces cerevisiae: implications to protein folding pathways in the cell. Mol Syst Biol. 2009;5 doi: 10.1038/msb.2009.26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Gille C., Goede A., Schlöetelburg C., Preissner R., Kloetzel P.M., Göbel U.B., Frömmel C. A comprehensive view on proteasomal sequences: implications for the evolution of the proteasome. J Mol Biol. 2003;326:1437–1448. doi: 10.1016/s0022-2836(02)01470-5. [DOI] [PubMed] [Google Scholar]
- 36.Wollenberg K., Swaffield J.C. Evolution of proteasomal ATPases. Mol Biol Evol. 2001;18:962–974. doi: 10.1093/oxfordjournals.molbev.a003897. [DOI] [PubMed] [Google Scholar]
- 37.Gupta R.S. Evolution of the chaperonin families (HSP60, HSP 10 and TCP-1) of proteins and the origin of eukaryotic cells. Mol Microbiol. 1995;15:1–11. doi: 10.1111/j.1365-2958.1995.tb02216.x. [DOI] [PubMed] [Google Scholar]
- 38.Lykke-Andersen S., Brodersen D.E., Jensen T.H. Origins and activities of the eukaryotic exosome. J Cell Sci. 2009;122:1487–1494. doi: 10.1242/jcs.047399. [DOI] [PubMed] [Google Scholar]
- 39.Veretnik S., Wills C., Youkharibache P., Valas R.E., Bourne P.E. Sm/Lsm genes provide a glimpse into the early evolution of the spliceosome. PLoS Comput Biol. 2009;5 doi: 10.1371/journal.pcbi.1000315. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Finnigan G.C., Hanson-Smith V., Stevens T.H., Thornton J.W. Evolution of increased complexity in a molecular machine. Nature. 2012;481:360–364. doi: 10.1038/nature10724. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41••.Johnston I.G., Dingle K., Greenbury S.F., Camargo C.Q., Doye J.P.K., Ahnert S.E., Louis A.A. Symmetry and simplicity spontaneously emerge from the algorithmic nature of evolution. Proc Natl Acad Sci USA. 2022;119 doi: 10.1073/pnas.2113883119. [DOI] [PMC free article] [PubMed] [Google Scholar]; The authors propose a non-adaptive mechanism explaining the prevalence of symmetric structures, which require less specific information to encode, and as such are more likely to appear than asymmetric assemblies through random mutations.
- 42.Kesmir C., van Noort V., de Boer R.J., Hogeweg P. Bioinformatic analysis of functional differences between the immunoproteasome and the constitutive proteasome. Immunogenetics. 2003;55:437–449. doi: 10.1007/s00251-003-0585-6. [DOI] [PubMed] [Google Scholar]
- 43••.Ascencio D., Diss G., Gagnon-Arsenault I., Dubé A.K., DeLuna A., Landry C.R. Expression attenuation as a mechanism of robustness against gene duplication. Proc Natl Acad Sci USA. 2021;118 doi: 10.1073/pnas.2014345118. [DOI] [PMC free article] [PubMed] [Google Scholar]; The authors measured the fitness effects of the duplication of several essential genes in the budding yeast to show that most duplications have small to undetectable effects on fitness. Further, attenuation of gene expression of the recently duplicated copies might be a mechanism to protect the multiprotein complexes from gene duplications.
- 44.Marcet-Houben M., Gabaldón T. Beyond the whole-genome duplication: phylogenetic evidence for an ancient interspecies hybridization in the baker’s yeast lineage. PLoS Biol. 2015;13 doi: 10.1371/journal.pbio.1002220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Oughtred R., Rust J., Chang C., Breitkreutz B.-J., Stark C., Willems A., Boucher L., Leung G., Kolas N., Zhang F., et al. The BioGRID database: a comprehensive biomedical resource of curated protein, genetic, and chemical interactions. Protein Sci. 2021;30:187–200. doi: 10.1002/pro.3978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Akiva E., Itzhaki Z., Margalit H. Built-in loops allow versatility in domain–domain interactions: lessons from self-interacting domains. Proc Natl Acad Sci USA. 2008;105:13292–13297. doi: 10.1073/pnas.0801207105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Hashimoto K., Panchenko A.R. Mechanisms of protein oligomerization, the critical role of insertions and deletions in maintaining different oligomeric states. Proc Natl Acad Sci USA. 2010;107:20352–20357. doi: 10.1073/pnas.1012999107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Gurel P.S., Kim L.Y., Ruijgrok P.V., Omabegho T., Bryant Z., Alushin G.M. Cryo-EM structures reveal specialization at the myosin VI-actin interface and a mechanism of force sensitivity. Elife. 2017;6 doi: 10.7554/eLife.31125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Knoll K.R., Eustermann S., Niebauer V., Oberbeckmann E., Stoehr G., Schall K., Tosi A., Schwarz M., Buchfellner A., Korber P., et al. The nuclear actin-containing Arp8 module is a linker DNA sensor driving INO80 chromatin remodeling. Nat Struct Mol Biol. 2018;25:823–832. doi: 10.1038/s41594-018-0115-8. [DOI] [PubMed] [Google Scholar]
- 50.Mistry J., Chuguransky S., Williams L., Qureshi M., Salazar G.A., Sonnhammer E.L.L., Tosatto S.C.E., Paladin L., Raj S., Richardson L.J., et al. Pfam: the protein families database in 2021. Nucleic Acids Res. 2021;49:D412–D419. doi: 10.1093/nar/gkaa913. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Thoden J.B., Holden H.M. The molecular architecture of galactose mutarotase/UDP-galactose 4-epimerase from Saccharomyces cerevisiae. J Biol Chem. 2005;280:21900–21907. doi: 10.1074/jbc.M502411200. [DOI] [PubMed] [Google Scholar]
- 52••.Nocedal I., Laub M.T. Ancestral reconstruction of duplicated signaling proteins reveals the evolution of signaling specificity. Elife. 2022;11 doi: 10.7554/eLife.77346. [DOI] [PMC free article] [PubMed] [Google Scholar]; The authors used the ancestral sequence reconstruction approach to experimentally demonstrate how interaction specificity evolves by fixation of point mutations after gene duplication in bacterial signaling pathways involving a two-component system.
- 53.Kortemme T., Joachimiak L.A., Bullock A.N., Schuler A.D., Stoddard B.L., Baker D. Computational redesign of protein-protein interaction specificity. Nat Struct Mol Biol. 2004;11:371–379. doi: 10.1038/nsmb749. [DOI] [PubMed] [Google Scholar]
- 54.Netzer R., Listov D., Lipsh R., Dym O., Albeck S., Knop O., Kleanthous C., Fleishman S.J. Ultrahigh specificity in a network of computationally designed protein-interaction pairs. Nat Commun. 2018;9 doi: 10.1038/s41467-018-07722-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Baker C.R., Hanson-Smith V., Johnson A.D. Following gene duplication, paralog interference constrains transcriptional circuit evolution. Science. 2013;342:104–108. doi: 10.1126/science.1240810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Kaltenegger E., Ober D. Paralogue interference affects the dynamics after gene duplication. Trends Plant Sci. 2015;20:814–821. doi: 10.1016/j.tplants.2015.10.003. [DOI] [PubMed] [Google Scholar]
- 57.Grandier-Vazeille X., Bathany K., Chaignepain S., Camougrand N., Manon S., Schmitter J.M. Yeast mitochondrial dehydrogenases are associated in a supramolecular complex. Biochemistry. 2001;40:9758–9769. doi: 10.1021/bi010277r. [DOI] [PubMed] [Google Scholar]
- 58.Huh W.-K., Falvo J.V., Gerke L.C., Carroll A.S., Howson R.W., Weissman J.S., O’Shea E.K. Global analysis of protein localization in budding yeast. Nature. 2003;425:686–691. doi: 10.1038/nature02026. [DOI] [PubMed] [Google Scholar]
- 59.Brevet A., Chen J., Lévêque F., Blanquet S., Plateau P. Comparison of the enzymatic properties of the two Escherichia coli lysyl-tRNA synthetase species. J Biol Chem. 1995;270:14439–14444. doi: 10.1074/jbc.270.24.14439. [DOI] [PubMed] [Google Scholar]
- 60.Musso R.E., Zabin I. Substrate specificity and kinetic studies on thiogalactoside transacetylase. Biochemistry. 1973;12:553–557. doi: 10.1021/bi00727a031. [DOI] [PubMed] [Google Scholar]
- 61.Lo Leggio L., Dal Degan F., Poulsen P., Andersen S.M., Larsen S. The structure and specificity of Escherichia coli maltose acetyltransferase give new insight into the LacA family of acyltransferases. Biochemistry. 2003;42:5225–5235. doi: 10.1021/bi0271446. [DOI] [PubMed] [Google Scholar]
- 62.Byrne K.P., Wolfe K.H. The Yeast Gene Order Browser: combining curated homology and syntenic context reveals gene fate in polyploid species. Genome Res. 2005;15:1456–1461. doi: 10.1101/gr.3672305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Ross J., Reid G.A., Dawes I.W. The nucleotide sequence of the LPD1 gene encoding lipoamide dehydrogenase in Saccharomyces cerevisiae: comparison between eukaryotic and prokaryotic sequences for related enzymes and identification of potential upstream control sites. J Gen Microbiol. 1988;134:1131–1139. doi: 10.1099/00221287-134-5-1131. [DOI] [PubMed] [Google Scholar]
- 64.Keyes B.E., Burke D.J. Irc15 Is a microtubule-associated protein that regulates microtubule dynamics in Saccharomyces cerevisiae. Curr Biol. 2009;19:472–478. doi: 10.1016/j.cub.2009.01.068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Lin A.-P., Demeler B., Minard K.I., Anderson S.L., Schirf V., Galaleldeen A., McAlister-Henn L. Construction and analyses of tetrameric forms of yeast NAD+-specific isocitrate dehydrogenase. Biochemistry. 2011;50:230–239. doi: 10.1021/bi101401h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Laurent M., Seydoux F.J., Dessen P. Allosteric regulation of yeast phosphofructokinase. Correlation between equilibrium binding, spectroscopic and kinetic data. J Biol Chem. 1979;254:7515–7520. [PubMed] [Google Scholar]
- 67.Breslow D.K., Collins S.R., Bodenmiller B., Aebersold R., Simons K., Shevchenko A., Ejsing C.S., Weissman J.S. Orm family proteins mediate sphingolipid homeostasis. Nature. 2010;463:1048–1053. doi: 10.1038/nature08787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Han G., Gupta S.D., Gable K., Bacikova D., Sengupta N., Somashekarappa N., Proia R.L., Harmon J.M., Dunn T.M. The ORMs interact with transmembrane domain 1 of Lcb1 and regulate serine palmitoyltransferase oligomerization, activity and localization. Biochim Biophys Acta Mol Cell Biol Lipids. 2019;1864:245–259. doi: 10.1016/j.bbalip.2018.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Marsh J.A., Teichmann S.A. Structure, dynamics, assembly, and evolution of protein complexes. Annu Rev Biochem. 2015;84:551–575. doi: 10.1146/annurev-biochem-060614-034142. [DOI] [PubMed] [Google Scholar]
- 70.Lynch M. The evolution of multimeric protein assemblages. Mol Biol Evol. 2012;29:1353–1366. doi: 10.1093/molbev/msr300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Conrad B., Antonarakis S.E. Gene duplication: a drive for phenotypic diversity and cause of human disease. Annu Rev Genom Hum Genet. 2007;8:17–35. doi: 10.1146/annurev.genom.8.021307.110233. [DOI] [PubMed] [Google Scholar]
- 72.Bailey J.A., Eichler E.E. Primate segmental duplications: crucibles of evolution, diversity and disease. Nat Rev Genet. 2006;7:552–564. doi: 10.1038/nrg1895. [DOI] [PubMed] [Google Scholar]
- 73.Verhaak R.G.W., Bafna V., Mischel P.S. Extrachromosomal oncogene amplification in tumour pathogenesis and evolution. Nat Rev Cancer. 2019;19:283–288. doi: 10.1038/s41568-019-0128-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Ronan A., Fagan K., Christie L., Conroy J., Nowak N.J., Turner G. Familial 4.3 Mb duplication of 21q22 sheds new light on the Down syndrome critical region. J Med Genet. 2007;44:448–451. doi: 10.1136/jmg.2006.047373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Moog U., Engelen J., Albrechts J., Hoorntje T., Hendrikse F., Schrander-Stumpel C. Alagille syndrome in a family with duplication 20p11. Clin Dysmorphol. 1996;5:279–288. [PubMed] [Google Scholar]
- 76.Khersonsky O., Tawfik D.S. Enzyme promiscuity: a mechanistic and evolutionary perspective. Annu Rev Biochem. 2010;79:471–505. doi: 10.1146/annurev-biochem-030409-143718. [DOI] [PubMed] [Google Scholar]
- 77.Nam H., Lewis N.E., Lerman J.A., Lee D.-H., Chang R.L., Kim D., Palsson B.O. Network context and selection in the evolution to enzyme specificity. Science. 2012;337:1101–1104. doi: 10.1126/science.1216861. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Lynch M. The evolution of genetic networks by non-adaptive processes. Nat Rev Genet. 2007;8:803–813. doi: 10.1038/nrg2192. [DOI] [PubMed] [Google Scholar]
- 79••.Hochberg G.K.A., Liu Y., Marklund E.G., Metzger B.P.H., Laganowsky A., Thornton J.W. A hydrophobic ratchet entrenches molecular complexes. Nature. 2020;588:503–508. doi: 10.1038/s41586-020-3021-2. [DOI] [PMC free article] [PubMed] [Google Scholar]; The authors show that evolutionary conservation of the oligomeric state can come from fixation of point-mutations that destabilize the monomeric state but not the multimeric state.
- 80.Plach M.G., Semmelmann F., Busch F., Busch M., Heizinger L., Wysocki V.H., Merkl R., Sterner R. Evolutionary diversification of protein-protein interactions by interface add-ons. Proc Natl Acad Sci USA. 2017;114:E8333–E8342. doi: 10.1073/pnas.1707335114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Jumper J., Evans R., Pritzel A., Green T., Figurnov M., Ronneberger O., Tunyasuvunakool K., Bates R., Žídek A., Potapenko A., et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596:583–589. doi: 10.1038/s41586-021-03819-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Baek M., DiMaio F., Anishchenko I., Dauparas J., Ovchinnikov S., Lee G.R., Wang J., Cong Q., Kinch L.N., Schaeffer R.D., et al. Accurate prediction of protein structures and interactions using a three-track neural network. Science. 2021;373:871–876. doi: 10.1126/science.abj8754. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Dey S, Ritchie DW, Levy ED. PDB-wide identification of biological assemblies from conserved quaternary structure geometry. Nat Methods. 2018;15:67–72. doi: 10.1038/nmeth.4510. [DOI] [PubMed] [Google Scholar]
- 84.Garcia Seisdedos H, Levin T, Shapira G, Freud S, Levy ED. Mutant libraries reveal negative design shielding proteins from supramolecular self-assembly and relocalization in cells. Proc Natl Acad Sci USA. 2022;119 doi: 10.1073/pnas.2101117119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Abrusán G, Marsh JA. Ligand-binding-site structure shapes allosteric signal transduction and the evolution of allostery in protein complexes. Mol Biol Evol. 2019;36:1711–1727. doi: 10.1093/molbev/msz093. [DOI] [PMC free article] [PubMed] [Google Scholar]